You are on page 1of 9

Bridging electromagnetic and carrier transport

calculations for three-dimensional modelling of


plasmonic solar cells
Xiaofeng Li,* Nicholas P. Hylton, Vincenzo Giannini, Kan-Hua Lee, Ned J. Ekins-
Daukes, and Stefan A. Maier
Blackett Laboratory, Department of Physics, Imperial College London, London SW7 2AZ, UK
*xfl_79@yahoo.com.cn

Abstract: We report three-dimensional modelling of plasmonic solar cells


in which electromagnetic simulation is directly linked to carrier transport
calculations. To date, descriptions of plasmonic solar cells have only
involved electromagnetic modelling without realistic assumptions about
carrier transport, and we found that this leads to considerable discrepancies
in behaviour particularly for devices based on materials with low carrier
mobility. Enhanced light absorption and improved electronic response
arising from plasmonic nanoparticle arrays on the solar cell surface are
observed, in good agreement with previous experiments. The complete
three-dimensional modelling provides a means to design plasmonic solar
cells accurately with a thorough understanding of the plasmonic interaction
with a photovoltaic device.
© 2011 Optical Society of America
OCIS codes: (040.5350) Photovoltaic; (240.6680) Surface plasmon; (290.1990) Diffusion.

References and links


1. H. A. Atwater, and A. Polman, “Plasmonics for improved photovoltaic devices,” Nat. Mater. 9(3), 205–213
(2010).
2. K. R. Catchpole, and A. Polman, “Plasmonic solar cells,” Opt. Express 16(26), 21793–21800 (2008).
3. V. Giannini, A. I. Fernández-Domínguez, Y. Sonnefraud, T. Roschuk, R. Fernández-García, and S. A. Maier,
“Controlling light localization and light-matter interactions with nanoplasmonics,” Small 6(22), 2498–2507
(2010).
4. V. E. Ferry, J. N. Munday, and H. A. Atwater, “Design considerations for plasmonic photovoltaics,” Adv. Mater.
22(43), 4794–4808 (2010).
5. S. A. Maier, M. L. Brongersma, P. G. Kik, S. Meltzer, A. A. G. Requicha, and H. A. Atwater, “Plasmonics – a
route to nanoscale optical devices,” Adv. Mater. 13(19), 1501–1505 (2001).
6. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, 2007).
7. E. M. Hicks, S. Zou, G. C. Schatz, K. G. Spears, R. P. Van Duyne, L. Gunnarsson, T. Rindzevicius, B. Kasemo,
and M. Käll, “Controlling plasmon line shapes through diffractive coupling in linear arrays of cylindrical
nanoparticles fabricated by electron beam lithography,” Nano Lett. 5(6), 1065–1070 (2005).
8. T. Shegai, V. D. Miljković, K. Bao, H. Xu, P. Nordlander, P. Johansson, and M. Käll, “Unidirectional broadband
light emission from supported plasmonic nanowires,” Nano Lett. 11(2), 706–711 (2011).
9. F. J. Beck, S. Mokkapati, A. Polman, and K. R. Catchpole, “Asymmetry in photocurrent enhancement by
plasmonic nanoparticle arrays located on the front or on the rear of solar cells,” Appl. Phys. Lett. 96(3), 033113
(2010).
10. S. Pillai, K. R. Catchpole, T. Trupke, and M. A. Green, “Surface plasmon enhanced silicon solar cells,” J. Appl.
Phys. 101(9), 093105 (2007).
11. P. Matheu, S. H. Lim, D. Derkacs, C. McPheeters, and E. T. Yu, “Metal and dielectric nanoparticle scattering for
improved optical absorption in photovoltaic devices,” Appl. Phys. Lett. 93(11), 113108 (2008).
12. K. Nakayama, K. Tanabe, and H. A. Atwater, “Plasmonic nanoparticle enhanced light absorption in GaAs solar
cells,” Appl. Phys. Lett. 93(12), 121904 (2008).
13. X. H. Chen, C. C. Zhao, L. Rothberg, and M. K. Ng, “Plasmon enhancement of bulk heterojunction organic
photovoltaic devices by electrode modification,” Appl. Phys. Lett. 93(12), 123302 (2008).
14. O. Stenzel, A. Stendal, K. Voigtsberger, and C. von Borczyskowski, “Enhancement of the photovoltaic
conversion efficiency of copper phthalocyanine thin film devices by incorporation of metal clusters,” Sol. Energy
Mater. Sol. Cells 37(3-4), 337–348 (1995).
15. T. L. Temple, G. D. K. Mahanama, H. S. Reehal, and D. M. Bagnall, “Influence of localized surface plasmon
excitation in silver nanoparticles on the performance of silicon solar cells,” Sol. Energy Mater. Sol. Cells 93(11),

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A888
1978–1985 (2009).
16. P. Spinelli, M. Hebbink, R. de Waele, L. Black, F. Lenzmann, and A. Polman, “Optical impedance matching
using coupled plasmonic nanoparticle arrays,” Nano Lett. 11(4), 1760–1765 (2011).
17. J. R. Nagel, and M. A. Scarpulla, “Enhanced absorption in optically thin solar cells by scattering from embedded
dielectric nanoparticles,” Opt. Express 18(Suppl 2), A139–A146 (2010).
18. R. A. Pala, J. White, E. Barnard, J. Liu, and M. L. Brongersma, “Design of plasmonic thin-film solar cells with
broadband absorption enhancements,” Adv. Mater. 21(34), 3504–3509 (2009).
19. W. Wang, S. M. Wu, K. Reinhardt, Y. L. Lu, and S. C. Chen, “Broadband light absorption enhancement in thin-
film silicon solar cells,” Nano Lett. 10(6), 2012–2018 (2010).
20. C. Hägglund, S. P. Apell, and B. Kasemo, “Maximized optical absorption in ultrathin films and its application to
plasmon-based two-dimensional photovoltaics,” Nano Lett. 10(8), 3135–3141 (2010).
21. J. N. Munday, and H. A. Atwater, “Large integrated absorption enhancement in plasmonic solar cells by
combining metallic gratings and antireflection coatings,” Nano Lett., (2010),
http://dx.doi.org/10.1021/nl101875t.
22. S. Mokkapati, F. J. Beck, A. Polman, and K. R. Catchpole, “Designing periodic arrays of metal nanoparticles for
light-trapping applications in solar cells,” Appl. Phys. Lett. 95(5), 053115 (2009).
23. J. R. Cole, and N. J. Halas, “Optimized plasmonic nanoparticle distributions for solar spectrum harvesting,”
Appl. Phys. Lett. 89(15), 153120 (2006).
24. M. D. Kelzenberg, S. W. Boettcher, J. A. Petykiewicz, D. B. Turner-Evans, M. C. Putnam, E. L. Warren, J. M.
Spurgeon, R. M. Briggs, N. S. Lewis, and H. A. Atwater, “Enhanced absorption and carrier collection in Si wire
arrays for photovoltaic applications,” Nat. Mater. 9(3), 239–244 (2010).
25. F. J. Beck, A. Polman, and K. R. Catchpole, “Tunable light trapping for solar cells using localized surface
plasmons,” J. Appl. Phys. 105(11), 114310 (2009).
26. J. Nelson, The Physics of Solar Cells (Imperial College Press, 2003).
27. Comsol Multiphysics, http://www.comsol.com/
28. G. A. Swartz, “Computer model of amorphous silicon solar cell,” J. Appl. Phys. 53(1), 712–719 (1982).
29. J. R. Lowney, and H. S. Bennett, “Majority and minority electron and hole mobilities in heavily doped GaAs,” J.
Appl. Phys. 69(10), 7102–7110 (1991).
30. H. S. Bennett, “Majority and minority electron and hole mobilities in heavily doped gallium aluminum arsenide,”
J. Appl. Phys. 80(7), 3844–3853 (1996).
31. D. E. Aspnes, S. M. Kelso, R. A. Logan, and R. Bhat, “Optical properties of AlxGa1x As,” J. Appl. Phys. 60(2),
754–767 (1986).
32. E. D. Palik, Handbook of Optical Constants of Solids, (Academic Press, 1985).
33. S. H. Lim, W. Mar, P. Matheu, D. Derkacs, and E. T. Yu, “Photocurrent spectroscopy of optical absorption
enhancement in silicon photodiodes via scattering from surface plasmon polaritons in gold nanoparticles,” J.
Appl. Phys. 101(10), 104309 (2007).
34. F. J. Beck, E. Verhagen, S. Mokkapati, A. Polman, and K. R. Catchpole, “Resonant SPP modes supported by
discrete metal nanoparticles on high-index substrates,” Opt. Express 19(Suppl 2), A146–A156 (2011).
35. G. W. Shu, W. C. Liao, C. L. Hsu, J. Y. Lee, I. J. Hsu, J. L. Shen, M. D. Yang, C. H. Wu, Y. C. Lee, and W. C.
Chou, “Enhanced conversion efficiency of GaAs solar cells using Ag nanoparticles,” Adv. Sci. Lett. 3, 368–372
(2010).
36. J. Meier, J. Spitznagel, U. Kroll, C. Bucher, S. Fay, T. Moriarty, and A. Shah, “Potential of amorphous and
microcrystalline silicon solar cells,” Thin Solid Films 451–452, 518–524 (2004).
37. A. J. Huber, F. Keilmann, J. Wittborn, J. Aizpurua, and R. Hillenbrand, “Terahertz near-field nanoscopy of
mobile carriers in single semiconductor nanodevices,” Nano Lett. 8(11), 3766–3770 (2008).

1. Introduction
The application of surface plasmons (SPs) in photovoltaic devices has recently attracted
considerable attention due to potential improvements to device absorption without increasing
the physical thickness of the photoactive medium [1–8]. Performance improvement has been
predicted through simulations and verified in experiments for solar cells (SCs) based on both
inorganic and organic materials [9–15] with properly designed plasmonic nanostructures [16–
25]. In order to comprehensively evaluate a SC, several crucial device parameters, e.g., short-
circuit photocurrent density (Jsc), current-voltage (I-V) curve, fill factor (FF), light-conversion
efficiency (η), etc. are required. These parameters can be readily achieved in photovoltaic
experiments [4,10,12,25]; however, most previous simulations of plasmonic SCs (PSCs)
consider only the optical absorption through solving Maxwell’s equations. Although improved
photocurrent has been predicted in several articles [17,21], ideal carrier transport (ICT) was
assumed, namely that each photo-generated electron-hole pair can successfully contribute to
the photocurrent. Under this condition the realistic solar cell behaviour and configuration
(e.g., doping profile, carrier mobility, carrier loss, etc.) are ignored. However, this approach is
not always accurate since in a real SC the photo-generated carriers will recombine (prior to
collection) under various loss mechanisms, such as surface and bulk carrier recombination or

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A889
short carrier diffusion [26]. The device performance is overestimated especially for those SCs
based on materials with low carrier mobility and short carrier lifetime, e.g., hydrogenated
amorphous silicon (α-Si:H). Hence, the exact electronic response must be considered in order
to obtain accurate I-V characteristics for a complete evaluation of the SCs. Therefore, an
accurate model of PSCs must consider both the electromagnetic (light absorption) and
electronic (carrier transport) properties together in a single simulation. Due to the strongly
localized confinement of SP waves, such work has to be performed in both frequency and all
spatial domains.
This paper presents exact three-dimensional (3D) simulations of PSCs that link both
optical absorption and carrier transport together. We have verified the need to perform the full
calculation of the electronic response of a SC by comparing the results of gallium arsenide
(GaAs) and a-Si:H SCs under a purely electromagnetic treatment and using our full
calculation. External quantum efficiency (EQE) and I-V characteristics before and after
plasmonic design were then calculated, in order to obtain complete information about the
performance improvement of SCs (e.g., Jsc, FF, η, etc). Finally, the spatial distributions of
power flow and carrier concentration profiles were extracted from the simulation.
2. Theory
Our work aims to realize a modular 2-step 3D simulation based on the finite element method
[27]. The first step is to perform an exact electromagnetic calculation (inherent in which are
all optical mechanisms, including any shading effect due to metallic nanoparticles on the
surface) based on 3D Maxwell’s equations in the frequency domain. A single unit cell was
used in conjunction with periodic Floquet-Bloch boundaries to represent the whole structure,
and the standard AM1.5G spectrum was introduced as the incident light source. Linearly
polarized light was used since the device is rotationally symmetric and insensitive to the light
polarization. The photo-generation rate extracted from the optical calculation is also used for
the exact carrier transport calculation in the second step, which is based on two sub-modules
simulating electron and hole transports, respectively, and one sub-module for electrostatic
potential (see the supplemental material). The 3D transport equations were solved under solar
injection and forward electric bias for the calculation of photocurrent and dark current.
Finally, information on SC performance, including EQE, Jsc, I-V curve, open-circuit voltage
(Voc), maximum output power density (Pmax), FF, η, etc. can all be obtained for the
optimisation of PSC designs.
The electronic response of solar cells (SCs) can be simulated accurately using three-
dimensional (3D) transport equations for electrons and holes. In order to develop a general
model capable of simulating various SCs, including those based on heterojunctions,
discontinuity of the intrinsic material parameters at heterojunction interfaces must be taken
into account. The simulation of homojunction SCs is then a specific case where the
discontinuity term is removed from the carrier transport equations. The generalized 3D carrier
transport equations (in steady state) used in our model are [26]
   k BT 
   Dn n  nn      ln N c    g ( x, y, z ,  )  U , (1)
  q q 

    E g k B T 
   D p p  p  p       ln N v    g ( x, y, z ,  )  U , (2)
  q q q  
where n (p) is the electron (hole) concentration, Dn = µnKBT/q (Dp = µpKBT/q) is the electron
(hole) diffusion coefficient, µn (µp) is the electron (hole) mobility, KB is Boltzmann’s
constant, T ( = 300 K) is the operating temperature, q is the electron charge, Ф is the
electrostatic potential, χ is the electron affinity, Eg is the material band gap, Nc (Nv) is the
effective conduction (valence) band density of states, g(x, y, z, λ) = α(x, y, z, λ)Ps(x, y, z, λ) is
the generation rate, α is the material absorption coefficient, Ps is the power flux calculated

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A890
from the electromagnetic model, and U is the rate of carrier loss due to Shockley-Read-Hall,
radiative, and Auger recombination.
The electrostatic potential Ф is determined by the charge profiles in the device according
to the 3D Poisson’s equation
q
2   n  p  C, (3)

where ε is the material permittivity, C = Nd – Na is the impurity concentration defined as the
sum of the concentrations of the ionized donors Nd and acceptors Na, including the signs of the
compensated charges.
The calculated spatial profiles of n and p are used to compute the frequency-dependent
photocurrent for electrons and holes, respectively [i.e., jn(x,y,z,λ) = qDnn and jp(x,y,z,λ) = –
qDpp]. The corresponding short-circuit current density jsc(λ) is then given by the averaged
photocurrent at the surface of the SCs
1 /2 /2
jsc ( ) 
2  
 / 2  / 2
jn ( x, y, L,  )  j p ( x, y, L,  ) dxdy, (4)

where L is the total thickness of the photoactive layers. The external quantum efficiency
(EQE) is then obtained by using EQE(λ) = jsc(λ)/[qbs(λ)], where bs is the solar incident flux.
The overall short-circuit current density Jsc can then be expressed by Jsc = ʃ jsc(λ)dλ.
Considering dark current Jd(V) and device resistances, the current-voltage (I-V) relation is
written as
V  J (V ) Rs
J (V )  J sc  J d (V )  , (5)
Rsh
where Rs and Rsh are series and shunt resistances, respectively, and Jd(V) is obtained by
applying forward electric bias into Eqs. (1) and (2), neglecting the generation terms. From the
I-V curve, parameters including open-circuit voltage Voc, maximum output power density
Pmax, fill factor [FF = Pmax/(JscVoc)], and light-conversion efficiency (η = Pmax/Psun, where Psun
is the overall incident light power density) can be obtained.
3. Simulation results
Table 1. Key Parameters Used in the Simulation [23,25–29].

Doping
µn µp
Layers τn τp
(cm–3)
(cm2/Vs) (cm2/Vs)
1 × 1018
GaAs SCs

window 100 10 1 ps 10 ns
emitter 4 × 1018
1100 80 1 ns 10 ns
base 2 × 1017
4000 80 1 ns 10 ns
BSF 2 × 1018
100 10 1 ns 1 ps
p region 4.6 × 102
1.3 × 1017 0.5 2 µs 0.34 µs
α-Si:H
SCs

i region 4.6 × 102 9.2 × 103 2 µs 2 µs


n region 4.3 × 10 16
1 9.2 × 103 1.7 µs 2 µs
BSF: back surface field. The doping concentration shown here for α-Si:H SCs is the ionised
donor/acceptor concentration; actual dopant levels can be much higher.

Listed in Table 1 [26,28–32] are the key parameters employed in the carrier transport
calculations. Additionally, the surface recombination coefficients for minority electrons and
holes are taken to be 1 × 106 (1 × 104) cm/s for GaAs (α-Si:H) SCs. Typical non-plasmonic
GaAs- and α-Si:H-based SCs [shown in Figs. 1(a) and 1(b) respectively] were first studied to
validate the necessity of performing the complete electronic calculation for a real SC. We
consider these two types of SCs since they are based on direct-band photoactive materials
with typically high and low carrier mobility, respectively, and can therefore test the two

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A891
modelling treatments under different conditions. Moreover, the large absorption coefficients
of the materials make them well-suited to thin-film PSC applications.
Simulation results are displayed in Figs. 1(c) – 1(f), where the results for “full model”
originate from the complete 3D calculation, while those for ICT were predicted by spatially
integrating and surface-averaging the carrier generation rate (multiplied by electron charge)
obtained from the optical calculation. Both current density and EQE of the SCs under solar
illumination were analyzed and a discrepancy between the estimation limited to
electromagnetic modelling and complete device modelling was observed. This discrepancy
relates to the device’s internal quantum efficiency (IQE), which can be directly approximated
by the ratio of the photocurrents derived from the exact calculation and the assumption of
ICT. For GaAs SCs, photocurrent and EQE are overestimated in the short-wavelength part of
the spectrum [Fig. 1(e)], while for α-Si:H SCs distinct overestimation was found over a much
broader spectral range [Fig. 1(f)]. We believe that the photocurrent decrease in real GaAs SCs
(compared to that estimated optically) primarily originates from surface recombination, which
is more pronounced at short wavelengths where the absorption coefficient increases
dramatically and a large fraction of the incident light is absorbed close to the surface.
However, α-Si:H SCs suffer from carrier loss due not only to recombination in the thin emitter
(including) but also very low carrier mobility. The latter occurs over a broad spectral range,
leading to a low IQE across much of the absorption spectrum. According to our calculation,
the exact value of Jsc under complete 3D calculation is 17.16 (9.15) mA/cm2 for the GaAs (α-
Si:H) SCs and that estimated from optical calculation is 17.93 (10.88) mA/cm 2. The
overestimated value is noticeable and impacts the PSC design (discussed later). It should be
noted that the accuracy of pure optical treatment strongly depends on the device
configuration; therefore, a robust model simulating both optical and electronic response is
necessary.

Fig. 1. (a) and (b): schematic diagrams of the considered GaAs- and α-Si:H SCs; (c) and (d):
current densities from solar incidence and those generated from the SCs; (e) and (f): EQE and
IQE response of the SCs. (a), (c) and (e) [(b), (d) and (f)] are for GaAs (α-Si:H) SCs. The
results obtained under ICT assumption are plotted with dashed curves. The observed peaks
between 550 nm and 800 nm for α-Si:H SCs are due to Fabry-Perot interference in the cavity.

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A892
Plasmonic effects can also be incorporated into our model for the simulation of PSC
structures [see Figs. 2(a) and 2(b)]. As an example, SCs were simulated with periodic arrays
of silver nanoparticles (with a diameter of 160 nm and 400 nm period) on the top surface. The
effect of plasmonic design on EQE response is displayed in Figs. 2(c) and 2(d), where both
EQE (i.e., photocurrent) degradation and enhancement can be observed. This is a typical
feature as reported in previous PSC experiments [12,25,33]. EQE decrease/increase in PSCs
was believed to be the result of the destructive/constructive interference between the scattered
field from the nanoparticles and that transmitted directly from the incidence, occurring on the
blue/red sides of the plasmonic resonance [33]. Alternatively, a recent explanation for this
phenomenon suggests that the photocurrent decrease at short wavelengths is due to the
resonant modes at the top of the particles, while the enhancement at long wavelengths is from
the modes localized at the Ag/substrate interface [34]. We have indeed observed similar
behaviour (not shown here) by checking the electric field distribution calculated from our
model. Note also that the spectra in Figs. 2(c) and 2(d) exhibit a sharp feature at a wavelength
of 400 nm, which is due to the first diffracted order arising from the periodicity of the
nanoparticles.
Under the proposed plasmonic design EQE enhancement was realized when λ > 455 nm
(447 nm) for GaAs (α-Si:H) PSCs, benefiting from the preferentially forward light scattering
and strong near-field confinement of SPs [1–4]. The plasmonic effect can be controlled
effectively by tuning the resonance of a single particle and the coupling strength of the
nanoparticles; therefore, the performance of PSCs can be optimized. On the other hand,
without considering carrier transport, as shown in Fig. 2, EQE is overestimated and the critical
wavelength distinguishing the photocurrent loss and gain regions is blue-shifted.

Fig. 2. Schematic diagram (a) [(b)], EQE response (c) [(d)] and I-V curves (e) [(f)] of GaAs [α-
Si:H] SCs before (dashed) and after (solid) the proposed plasmonic design with 160nm-
diameter silver particles under period of 400 nm. Solid and dot curves are from complete
calculation and optical estimation under ICT assumption, respectively. Power densities are also
plotted in (e) and (f) so that the detailed information about Jsc, Voc, Pmax, FF, and η can be
obtained. The performance enhancement is observed from these figures and the extracted
performance parameters are listed in Table 2.

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A893
In order to acquire the I-V information, the carrier transport module is used again to
calculate the dark current response of the SCs under forwardly applied voltages. This was
realized in our model by controlling the boundary conditions of the electrostatic potential sub-
module [26]; the results are plotted in Figs. 2(e) and 2(f), respectively. In addition, output
power densities have also been plotted so that Pmax, FF, and η could be determined. From
these figures, the device performance parameters under various conditions were extracted and
listed in Table 2, which shows that, with the incorporation of plasmonic nanostructures,
improvement ratios in Jsc, Pmax, and η of more than 18% for GaAs SCs and over 40% for α-
Si:H SCs could be achieved. As expected, over-predicted performance under ICT assumption
was observed. The difference between these two methods is especially distinct for α-Si:H SCs
as shown in Fig. 2(f) and Table 2.
Table 2. Performance Comparison under Various System Configurations

Jsc Voc Pmax


FF η
(mA/cm2) (V) (mW/cm2)
no SP 17.16 1.028 14.34 81.3% 14.9%
GaAs SCs

PSC (full model) 20.28 1.034 17.10 81.6% 17.8%


(enhancement, %) (18.2%) (0.58%) (19.3%) (0.37%) (19.5%)
PSC (ICT) 21.00 1.035 17.73 81.6% 18.4%
(enhancement, %) (22.4%) (0.68%) (23.6%) (0.37%) (23.5%)
no SP 9.15 0.94 7.18 83.5% 7.5%
α-Si:H SCs

PSC (full model) 12.78 0.95 10.20 84.0% 10.6%


(enhancement, %) (39.7%) (1.1%) (42.1%) (0.6%) (41.3%)
PSC (ICT) 15.12 0.96 12.14 83.6% 12.16%
(enhancement, %) (65.3%) (2.1%) (69.1%) (0.12%) (68.0%)

However, neither Voc nor FF exhibit noticeable increases [11,35]. Note that the FF shown
here is relatively high, since ideal series (i.e., 0 Ωcm2) and shunt resistances (i.e.,  Ωcm2)
were used for simplicity. Realistic resistance values could be easily integrated into our model
[see Eq. (5)], resulting in a more credible FF. Note also that the incorporation of metallic
components on the incident surface may lead to an associated reduction in resistance, hence
increasing the FF [12,26]. This effect can be included by using appropriate device resistances
that reflect the effect of a metallic nanoparticle coating. We also emphasize that the scope of
our work is to obtain a full 3D model that considers both the electromagnetic and carrier
transport response, and that the optimization of PSCs is beyond the scope of this paper; hence
the device performance shown herein may not be maximized compared to some previous
reports (e.g., for α-Si:H SCs reported in [36]).
We would like to indicate that there also exists a compromise between a full 3D model
and the ICT case by using only low-dimensional electronic model, i.e., based on the
assumption that the IQE of the original SC shows no noticeable change after the introduction
of plasmonic nanoparticles. In this compromise the original IQE from a 1D electronic
calculation and the field profile of the PSCs from our 3D electromagnetic calculation are used
to predict the final electronic output. Using this method, Jsc values for the considered GaAs
and α-Si:H PSCs were recalculated and found to be 20.25 and 13.00 mA/cm 2, respectively,
showing an obvious improvement from the ICT case. However, it is unclear if such a
compromise remains valid for all cases (different materials and system setups or strong
plasmonic effects). For example, if the solar cells are under multi-sun illumination or a higher
nanoparticle concentration is employed, the carrier profiles will be modified more strongly by
plasmonic design leading to significant modification of the IQE. Furthermore, obtaining
detailed spatial information of the electronic parameters is also an important issue, which can
only be obtained from a complete 3D simulation.
Therefore, we now turn our attention to the benefits of our 3D simulation in enabling the
extraction of spatial information on optical and electronic parameters. Shown in Fig. 3 are the
power flows in the α-Si:H PSCs (GaAs PSCs show qualitatively same behaviour) working in
the photocurrent loss [λ = 350 nm, Figs. 3(a1) and 3(a2)] and gain [λ = 500 nm, Figs. 3(b1) –
3(b3)] regions, respectively. The images shown in Fig. 3(b3) correspond to the various z

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A894
positions shown in Fig. 3(b1). The figure clearly shows that the uniform distribution in the xy
plane and exponential decay along the z direction in a conventional non-plasmonic SC have
been strongly modified due to the presence of plasmonic nanostructures. Power confined in
the active layers can be either decreased under short-wavelength incidence [see Figs. 3(a1)
and 3(a2)] or increased under long-wavelength incidence [see Figs. 3(b1) – 3(b3)]. In the
photocurrent gain region, the incident light can undergo strong forward scattering and
confinement in the photoactive regions. This is especially beneficial to increase the effective
optical absorption length so that photocurrent can be generated more efficiently.

Fig. 3. Power flow distributions in the plasmonic α-Si:H SCs with 160nm-diameter silver
nanoparticles decorated above. In (a1) and (a2), λ = 350 nm (in photocurrent loss region) and in
(b1), (b2), and (b3) λ = 500 nm (in photocurrent gain region). In (b3), field polarizations are
also given.

Figure 4 illustrates the calculated distributions of carrier generation rate and electron and
hole concentrations inside the active layers of α-Si:H SCs working at λ = 500 nm [in the
region of photocurrent gain as shown in Fig. 2(b)]. It is clear that the spatially modulated
incident wave shown in Fig. 3 reshapes the carrier generation profile as well as modifies the
carrier concentration distribution. However, the overall carrier concentration profiles are not
significantly changed after solar injection (not shown here). This is firstly because the photo-
generated carrier concentration under one-sun illumination is far less than the background;
secondly, the spatial non-uniformity will be alleviated with carrier transporting inside the
active layers.
The transverse dependence of the stabilized carrier profiles arising from the plasmonic
nanostructures can be best seen on a logarithmic scale, as shown by Figs. 4(b1) – 4(c2). In
addition, spatial information of other electronic properties (e.g. electrostatic potential,
electrostatic field, carrier recombination rate, etc.) can also be obtained from our simulation.
Here, we would like to emphasize that the ability to obtain detailed carrier concentration
distributions inside the device is an important and experimentally accessible value. For
example, a terahertz near-field microscopy technique was recently reported [37], where the
authors demonstrated that the technique is capable of measuring carrier densities close to the
surface of a material. They showed that regions with metals or highly conductive

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A895
semiconductors appear brightest in the THz image while semiconductors with low doping
levels appear darker. Our model can provide a powerful theoretical tool to aid interpretation or
corroborate the results of the terahertz near-field microscopy technique. Our model is also
able to calculate the 3D carrier distribution throughout the materials, not just at the surface but
internally. Since carrier recombination proceeds via different channels as various powers of
carrier density, it is critical to account for local variation in carrier density to accurately
predict photovoltaic device performance.

Fig. 4. Distributions of carrier generation rate [(a1) & (a2)], electron concentration [(b1) &
(b2)], and hole concentration [(c1) & (c2)] inside the active layers of α-Si:H PSCs working at λ
= 500 nm (in photocurrent gain region). Active layer configuration has been shown in Figs.
1(b) and 3(a1).

4. Conclusion
In summary, PSCs have been simulated using our 3D model, which calculates both the optical
and electronic responses of a solar cell structure. Besides the extensively investigated optical
properties, it also provides 1) a realistic measurement of the electronic performance
enhancements arising from the presence of metallic nanoparticles and 2) spatial distributions
of electronic parameters inside the device. Comparison between complete device simulation
and optical estimation shows that carrier transport needs to be considered in a realistic SC
model. Typical findings agreeing with PSC experiments and performance improvement with
the incorporation of plasmonic design are verified. Since this model limits neither plasmonic
nanostructure geometry nor SC configuration, it can be widely applied for the modelling of
PSCs and conventional non-plasmonic SCs in 3D. Minor modifications to the model would
enable light-emitting diodes and photodetectors to be modelled in a similar way.
Acknowledgments
This work was supported by EU FP7 project “PRIMA” - 248154.

#145911 - $15.00 USD Received 13 Apr 2011; revised 11 May 2011; accepted 13 May 2011; published 30 Jun 2011
(C) 2011 OSA 4 July 2011 / Vol. 19, No. S4 / OPTICS EXPRESS A896

You might also like